You are on page 1of 11

Acta Bot. Croat.

79 (1), 15–25, 2020 CODEN: ABCRA 25


DOI: 10.37427/botcro-2020-002 ISSN 0365-0588
eISSN 1847-8476

Preliminary characterization of the Quercus


pubescens complex in southern Italy using molecular
markers
Romeo Di Pietro1, Piera Di Marzio2, Gaby Antonecchia3, Antonio Luca Conte2, Paola Fortini2*
1
 epartment of Planning, Design, Architecture technology (PDTA), Herbarium Flaminio, Sapienza University of Rome,
D
Rome, Italy
2
Department of Bioscience and Territory, University of Molise, Pesche (IS), Italy
3
Ministry of Education, University and Research (MIUR), Rome, Italy

Abstract – Quercus pubescens s.l. is a group of taxonomically intricate and highly debated deciduous white oaks
widely distributed in southern Europe. The Apulia region occupies the south-easternmost part of the Italian Pen-
insula; the land-use pattern is based on extensive agricultural systems and only 10% is covered by forests that are
mainly composed of oak woods. It is the region in Italy showing the highest number of oak species, among which
four putative species of the Quercus pubescens group, have been reported in floras and checklists with uncertain
taxonomic value because of the overlapping of diagnostic characters. In this paper, we carried out a molecular
analysis on natural populations of Q. pubescens s.l. distributed throughout the Apulia region. Individuals from 24
pubescent oak populations were sampled and each tree was genotyped at 11 polymorphic microsatellite markers.
Overall, the average expected heterozygosity (He) was 0.629, and the allelic richness (Ar) ranged between 2.130
and 7.187. No differentiation was observed among the populations investigated, and the genetic differentiation
coefficient (FST) was 0.036. Gene flow among populations was found to be relatively high (Nm = 6.664). From a
taxonomic point of view, the possibility of the coexistence of more than one species among the Apulian pubes-
cent oaks reported in the taxonomic and syntaxonomic literature is not supported by the results of this molecular
analysis.

Key words: genetic diversity; Italy; population structure; Quercus pubescens; SSR; taxonomy

Introduction
The genus Quercus L. subgenus Quercus has a wide dis- dalechampii Ten., Q. leptobalana Guss., Q. ichnusae Mossa,
tribution in the Northern Hemisphere, especially in central Bacch. et Brullo, Q. humilis Mill., Q. virgiliana Ten.) which
and southern Europe, where Quercus species form important are all considered Steno-Mediterranean vicariant species of
forest communities in the Temperate and the Mediterranean Q. pubescens Willd.
bioclimatic regions (Nixon 1993, Govaerts and Frodin 1998). Apulia is the easternmost Italian administrative region
Italy is the country showing the highest oak species di- and all Italian oak species, both evergreen and deciduous,
versity in Europe, although there is still incomplete agree- occur in this region with the exception of Q. petraea (Matt.)
ment on the exact number of oak taxa occurring in the ter- Liebl. Despite this richness in oak species and a wide poten-
ritory. Most of the taxonomic uncertainty regarding the tial range for the oak woodlands, only 10% of the Apulian
Italian (and European) oaks concerns the so-called white territory is covered by forests. The Q. pubescens s.l. forests
oaks and in particular the pubescent oaks (Quercus pube- are the most widespread and can be found in the flooded de-
scens s.l. subgen. Quercus Sect. Quercus). The most recent pressions of the plain, in the limestone plateaus and on the
Italian flora and checklists (Pignatti et al. 2017, Bartolucci et sub-montane rocky slopes.
al. 2018) report various pubescent oaks in southern Italy (Q. Molecular marker studies have not yet been published
amplifolia Guss., Q. apennina Lam., Q. congesta C. Presl, Q. for the genus Quercus in the Apulian region. However, the

* Corresponding author e-mail: fortini@unimol.it

ACTA BOT. CROAT. 79 (1), 2020 15


DI PIETRO R, DI MARZIO P, ANTONECCHIA G, CONTE AL, FORTINI P

evaluation of genetic variation can be of great importance The bedrock of the mountainous systems is mainly com-
for the conservation and management of forest ecosystems, posed of Cretaceous limestone and Paleocenic calcarenites.
especially in areas particularly vulnerable to climatic change Marls occur in the lower parts of the Daunian sub-Apennine
(Peñuelas et al. 2017) and in forests whose original extent has and in the Tavoliere plain together with conglomerates and
been significantly reduced. This is the case of the oak forests Pleistocenic and Holocenic sand deposits. The bioclimate is
of the Apulia region. In fact, the dry Mediterranean climate Thermo-Mediterranean and Meso-Mediterranean with the
that characterizes this region places it at risk of desertifica- Temperate region occurring only in the sub-and lower-mon-
tion if the climate should experience a further increase in tane belts. The rainiest areas are the Gargano, the Daunian
aridity. Moreover, the widespread practice of extensive ag- sub-Apennine and the south-eastern part of the Salento Pen-
riculture (wheat, olive and vine) during the last century led insula (slightly over 800 mm year–1). Annual precipitation
to the destruction of most of the natural forest resources values on average of less than 500 mm year–1 are recorded
(Biondi et al. 2010). in the western side to the Murgian Plateau and in the Tavo-
Several reports on the population genetics of the white liere plateau whereas the remaining portion of the territory
oaks Q. robur L., Q. petraea, Q. pubescens have been pub- exhibits an average annual rainfall of between 500 and 700
lished for south-eastern Europe in the last two decades mm year–1 (Blasi and Michetti 2007, Cotecchia et al. 2014).
(Franjić et al. 2006, Curtu et al. 2007a, Jerše and Batič 2007, As far as the climax vegetation is concerned almost the en-
Slade et al. 2008, Curtu et al. 2009, Ballian et al. 2010, Enes- tire Apulian territory would potentially be covered by oak
cu et al. 2013, Gailing et al. 2013). In Italy, only few data are forests except for some scattered coastal and subcoastal ar-
available. Fineschi et al. (2002) and Fineschi and Vendramin eas dynamically linked to the Pinus halepensis Mill. woods
(2004) presented a study on the genetic diversity of the Ital- and maquis, and the inner part of the Gargano promontory
ian white oaks analyzing chloroplast DNA. Subsequently, where the wide “Foresta umbra” beech wood occurs (Biondi
several studies that compared leaf morphology and molecu- et al. 2004, 2010).
lar data on some white oaks populations of Central Italy have
been published (Fortini et al. 2009, 2013, 2015). Population sample information
In the most recent papers published on the Apulia forests Individual mature oak trees were randomly selected in
(Biondi et al. 2004, Di Pietro and Misano 2009) the follow- each population trying to maintain a distance of at least 20
ing pubescent oaks were reported in the phytosociological meters from each other. Leaves were randomly collected
tables: Q. pubescens, Q. virgiliana, Q. amplifolia and Q. dale- from 379 individuals within 24 populations (subdivided
champii. A recent study (Di Pietro et al. 2016) statistically in squares 50 × 50 m), which covered approximately uni-
analyzed 25 morphological characters of leaves and fruits formly the physiographic units of the Apulia region (Fig.
within 24 pubescent oak populations in the Apulia region, 1, Tab. 1, On-line Suppl. Tab. 1). Both the individuals and
in order to identify possible diagnostic traits useful for taxo- the populations they belong to are the same on which the
nomic differentiation. The results provided no evidence for
morphological analysis of leaves and acorns was performed
differentiation among tree individuals based on compari-
in the paper by Di Pietro et al. (2016). The collected oak
sons of morphological traits. In the present study plant ma-
specimens were dried on silica gel, and stored at room tem-
terial from the same pubescent oak populations was used to
perature until analysis. Voucher specimens for each indi-
analyze their genetic diversity, structure and gene flow. The
vidual sampled were stored at the Herbarium of the Uni-
aims of this study are twofold. First, to assess whether any
versity of Molise.
group of genetic structure could be identified among oak
individuals and/or populations, notwithstanding the mor-
DNA extraction, EST-SSR amplification
phological and taxonomical uniformity highlighted in the
previous morphological work (Di Pietro et al. 2016). Second, Total genomic DNA was extracted from 0.5 g of dried
to investigate the pattern of genetic diversity in the Apulian leaves from all samples using spin columns of “Invisorb®
pubescent oaks in order to provide useful information for Spin Plant Mini Kit” and following the protocol of the man-
further studies. ufacturer.
Eleven microsatellite loci or simple sequence repeats
Material and methods (EST-SSRs), developed by Durand et al. (2010) (PIE020,
PIE102, PIE152, PIE215, PIE223, PIE227, PIE239, PIE242,
Study area PIE243, PIE267, PIE271) and polymorphic in white oaks
The Apulia region is located in the south-eastern part of (Guichoux et al. 2011, Antonecchia et al. 2015) were se-
the Italian Peninsula where it is largely open to the Adriatic lected. Polymerase chain reaction (PCR) amplification was
and Ionian Seas. It is slightly sloping, with more than 60% performed in a single multiplex reaction in a DNA Engine
of the territory occurring below 200 m a.s.l. Five physio- Tetrad (MJ Research Bio-Rad) thermocycler and PCR prod-
graphic units can be distinguished in the Apulia region and ucts were run on an ABI 3730xl capillary sequencer (Applied
these are the Daunian sub-Apennine (1150 m), the Garga- Biosystems) using Genescan 600 LIZ internal size standard
no promontory (1080 m), the Murgian Plateau (680 m), the (Applied Biosystems). Alleles were scored using STR and
Salento Peninsula and the large plain Tavoliere delle Puglie. software version 2.3.106 (Toonen and Hughes 2001) and al-
16 ACTA BOT. CROAT. 79 (1), 2020
QUERCUS PUBESCENS COMPLEX IN SOUTHERN ITALY

Fig. 1. Locations of the 24 oak collection stands.

Tab. 1. Quercus species composition of the 24 oak populations (according to Biondi et al. 2004, Biondi and Guerra 2008, Di Pietro and
Misano 2009) with number of genotyped individuals and coordinates.
Population Analyzed Genotyped
Latitude (N) Longitude (E) Quercus species
code individuals individuals
Pop01 16 16 4524805 650840 Q. pubescens
Pop02 16 16 4557220 526723 Q. pubescens
Pop03 16 16 4624669 567854 Q. pubescens
Pop04 16 16 4624043 561581 Q. virgiliana, Q. dalechampii, Q. pubescens
Pop05 16 15 4562416 534533 Q. virgiliana, Q. dalechampii, Q. pubescens
Pop06 16 15 4564231 529921 Q. pubescens
Pop07 16 16 4558496 530175 Q. pubescens
Pop08 16 16 4512469 618357 Q. pubescens
Pop09 18 16 4497035 647716 Q. pubescens
Pop10 16 16 4582747 553553 Q. pubescens, Q. dalechampii, Q. virgiliana
Pop11 16 16 4551101 622237 Q. virgiliana, Q. dalechampii, Q. amplifolia
Pop12 15 15 4504016 656599 Q. pubescens
Pop13 12 12 4491270 658881 Q. pubescens, Q. dalechampii, Q. virgiliana
Pop14 16 16 4446957 278106 Q. virgiliana, Q. amplifolia, Q. dalechampii
Pop15 15 15 4514201 647317 Q. pubescens
Pop16 16 11 4502657 686693 Q. pubescens
Pop17 16 13 4534971 632507 Q. pubescens, Q. dalechampii, Q. virgiliana
Pop18 16 14 4523011 649110 Q. dalechampii, Q. virgiliana, Q. pubescens
Pop19 16 14 4499728 654902 Q. pubescens
Pop20 17 14 4635503 582284 Q. pubescens, Q. dalechampii, Q. virgiliana
Pop21 16 16 4504890 744270 Q. virgiliana, Q. dalechampii, Q. pubescens
Pop22 14 11 4518821 650238 Q. virgiliana, Q. dalechampii, Q. pubescens
Pop23 16 6 4540867 597285 Q. pubescens
Pop24 16 7 4635677 587298 Q. pubescens

lele binning was performed as described in Guichoux et al. populations. EST-SSR amplification was unsuccessful for 41
(2011). MICROCHECKER (Van Oosterhout et al. 2004) was of the 379 individuals sampled. Two populations (Pop23 and
used to check for the presence of null alleles and to check for Pop24) were not included in the statistical analyses because
potential problems related to allele dropout with 1000 ran- of the low number of amplified individuals.
domizations and a 95% confidence interval. The basic statistics for each locus: number of alleles
(NA), number of individuals (N), observed heterozygosity
Data analysis (Ho), expected heterozygosity (He), and number of private
Polymorphic EST-SSR markers were used to analyze the alleles (PA) were computed using the statistical program Ge-
genetic diversity of 338 Q. pubescens s.l. individuals in 24 nAlEx version 6.503 (Peakall and Smouse 2012).
ACTA BOT. CROAT. 79 (1), 2020 17
DI PIETRO R, DI MARZIO P, ANTONECCHIA G, CONTE AL, FORTINI P

Polymorphic information content (PIC), deviation from value were done to test the consistency of the results. Each
Hardy-Weinberg equilibrium (HW), and null allele frequen- (K) value was tested from K = 1 to K = 10. The optimum K
cy (F(Null)) were computed for each locus using Cervus value was predicted using the web-based software STRUC-
software version 3.0.7 (Kalinowski et al. 2007). TURE HARVESTER web version 0.6.94 which implements
The basic statistics for each population: number of dif- the Evanno method (Earl and vonHoldt 2012). Individu-
ferent alleles, number of effective alleles, allelic richness, ob- als with probabilities above 0.80 (Q ≥ 0.800) were assigned
served heterozygosity, expected heterozygosity, index of fix- as putative purebred while all the other individuals were
ation (FIS) calculated as F = 1 – (Ho / He) and number of assigned as “putative hybrids” although the term “hybrid”
private alleles, were computed using Arlequin software ver- could be a little misleading since no plant material of pos-
sion 3.5.2.2 (Excoffier and Lischer 2010). Allelic richness per sible parental species was collected. For this reason we have
locus and population were computed using FSTAT version opted for using the much more neutral term “off-type” in
2.9.4 (Goudet 2003). the rest of the text.
BOTTLENECK software version 1.2 (Cornuet and Lui-
Analysis of molecular variance (AMOVA) was per-
kart 1996) was used to detect the likelihood of a bottleneck
formed using GenAlEx in order to partition the total mi-
effect. Heterozygosity excesses were displayed, identified
crosatellite diversity, among and within populations, and
based on the estimates of multilocus genotypes calculated
within the individuals of each Q. pubescens s.l. population,
by using the Wilcoxon signed rank tests and evaluating de-
based on 11 EST-SSR markers loci. The variance compo-
partures from the ratio 1:1 deficiency/excess (Cornuet and
nents were statistically tested by non-parametric random-
Luikart 1996, Luikart and Cornuet 1998, Piry et al. 1999).
ization tests of significance based on 1000 bootstraps. Gene
A two-phase mutation model (TPM) and a stepwise muta-
flow among populations was estimated using the indirect
tion model (SMM) for Wilcoxon signed-rank tests were used
method based on the number of migrants per generation considering a 90% SMM proportion and TPM with a vari-
(Nm) using the formula, Nm = 0.25 (1 − FST) / FST (Slatkin ance of 10, and 1000 repeats.
and Barton 1989). To investigate population differentiations,
pairwise FST between all pairs of populations was computed,
FST was calculated according to Weir and Cockerham (1984) Results
in program FSTAT 2.9.4. SSR polymorphism and genetic diversity
Principal coordinate analysis (PCoA) with data stan- All the loci were polymorphic (Pj = q ≤ 0.99). The total
dardization based on Nei’s genetic distance was performed number of alleles identified was 100 out for the 325 individu-
using GenAlEx. als analyzed. The allele fragment sizes (Tab. 2) matched with
Bayesian clustering analysis, using STRUCTURE soft- the relative SSR reference size, as described in Guichoux et
ware version 2.3.4 (Pritchard et al. 2000, 2010), with popu- al. (2011). The number of alleles per locus (k) ranged be-
lation IDs as Sampling Location Information, has been car- tween 4 and 17, with an average of 11 alleles. The PIC values
ried out. We opted for the admixture with correlated allele varied from 0.233 (PIE227) to 0.879 (PIE152). According to
frequencies model between populations. We used a burn-in (Hildebrand et al. 1992) eight EST-SSR loci were found to
period of 25,000 and a number of MCMC of 100,000 cy- be informative (PIC ≥ 0.5), whereas two were found to be
cles. Ten runs for each K (K = number of possible clusters) slightly informative (PIC < 0.4). The observed heterozygos-

Tab. 2. Allelic diversity of eleven microsatellite loci scored in 325 individuals of Quercus pubescens s.l.; S – size range fragments (bp);
NA – Number of alleles at the locus, Ar – Allelic richness per locus, N – Number of individuals typed at the locus, Ho – Observed
heterozygosity, He – Expected heterozygosity, PIC – polymorphic information content, HW – Significance of deviation from Hardy-
Weinberg equilibrium (Hardy-Weinberg equilibrium test chi-square value, p value and significance with Bonferroni correction: ns =
not significant, * = P < 0.05; ** = P < 0.01; *** = P < 0.001), F(Null) – Null allele frequency estimate, FIS – Fixation index (* = P < 0.05).
Locus S NA Ar N Ho He PIC HW F(Null) FIS
PIE020 97–109 7 3.386 312 0.583 0.585 0.537 35.434*** –0.026 –0.067
PIE102 139–167 13 5.367 321 0.523 0.750 0.726 87.243 *** 0.187 0.279*
PIE152 228–260 17 7.187 314 0.834 0.890 0.879 3.756 ns 0.032 0.043*
PIE215 185–218 12 6.295 281 0.384 0.707 0.673 193.668*** 0.311 0.417*
PIE223 197–234 11 5.980 318 0.811 0.822 0.803 0.525 ns 0.006 –0.002
PIE227 156–165 4 3.387 323 0.223 0.251 0.233 3.776 ns 0.060 0.095*
PIE239 69–93 9 4.689 302 0.248 0.435 0.420 56.089*** 0.274 0.399*
PIE242 102–132 14 5.924 318 0.748 0.841 0.821 16.677ns 0.058 0.093*
PIE243 204–224 10 4.392 282 0.688 0.683 0.649 37.002*** –0.029 –0.038
PIE267 84–106 12 4.029 313 0.425 0.667 0.622 82.591*** 0.227 0.334*
PIE271 182–204 12 2.130 311 0.949 0.843 0.823 37.040*** –0.063 –0.153
Mean – 11 – 308.64 0.583 0.679 0.653 – 0.094 –
Standard error – 1.053 – 4.386 0.074 0.058 0.059 – 0.040 –

18 ACTA BOT. CROAT. 79 (1), 2020


QUERCUS PUBESCENS COMPLEX IN SOUTHERN ITALY

ity (Ho) for microsatellite loci ranged from 0.223 (PIE227) ent observed alleles (Na) and effective alleles (Ne) averaged
to 0.949 (PIE271), and the average was 0.583. A high level of across all loci ranged from 4.727 (Pop22) to 7.182 (Pop09)
Ho was detected in seven EST-SSR loci ranging from 0.523 and 2.559 (Pop22) to 4.363 (Pop20), respectively. Nine pri-
to 0.949, whereas four loci detected low Ho with values rang- vate alleles (PA) was found and these are distributed within
ing from 0.223 to 0.425. The expected heterozygosity (He) ten individuals in turn distributed in eight oak populations
ranged from 0.251 to 0.890 with an average value of 0.679 (two other private alleles were found in Pop24; On-line Sup-
and was higher than Ho for all the microsatellite loci, ex- pl. Tab. 2). The allelic richness (Ar), ranged between 3.865
cept for PIE243 and PIE271. The Hardy-Weinberg exact test (Pop3) and 5.171 (Pop20). The average observed heterozy-
for all populations (HW) revealed that seven loci (PIE020, gosity (aHo) ranged from 0.487 (Pop03) to 0.665 (Pop11) with
PIE102, PIE 215, PIE239, PIE243, PIE267, PIE271) exhib- mean value of 0.583. The average expected heterozygosity
ited significant deviation from Hardy–Weinberg equilibri- (aHe) ranged from 0.510 (Pop03) to 0.709 (Pop09) with mean
um (P<0.001) whereas four loci (PIE152, PIE223, PIE227, value of 0.629. Overall, the observed heterozygosity showed
PIE242) did not show significant departures from Hardy- mostly slightly lower values than the expected heterozygosity.
Weinberg equilibrium. The mean fixation index (FIS) values are close to zero for all
The indexes of genetic diversity of the 22 Q. pubescens s.l. populations (ranging from 0.058 in Pop11 to 0.188 in Pop21)
populations are summarized in Tab. 3. The number of differ- that are the values expected under a random mating.

Population genetic structure and gene flow


Tab. 3. Genetic diversity parameters for the Quercus pubescens The overall population differentiation degree is reported
s.l. populations (Pop) analyzed through eleven microsatellite loci. in Tab. 4. AMOVA analysis showed that 3.6% (P < 0.001) of
n – number of individuals; Na – average number of different al-
leles; Ne – average number of effective alleles; Ar – allelic richness; the genetic variations was among populations while 96.4%
aHe – average expected heterozygosity; aHo – average observed was within populations. The majority of molecular variance
heterozygosity; FIS – average of fixation index tested by 1023 per- was partitioned within individuals (78.5%, P < 0.01) while
mutation of gene copies between individuals within population
17.9% (P < 0.01) of variance was found to occur among in-
(* = P < 0.05); PA – number of private alleles.
dividuals. The pairwise FST matrix between populations is
Pop n Na Ne Ar aHo aHe FIS PA shown in On-line Suppl. Tab. 3; eleven population pairs ex-
Pop01 16 6.182 3.641 4.642 0.581 0.604 0.035 0 hibited negative values but all of them showed no significant
Pop02 16 5.091 3.353 4.064 0.498 0.584 0.125* 0 p values (all largely > 0.05). The differentiation coefficient
Pop03 16 4.818 2.968 3.865 0.487 0.510 0.029 0 between population pairs having significant P values showed
Pop04 16 6.364 3.684 4.605 0.572 0.638 0.094* 0
that Pop11 (Incoronata-Tavoliere) and Pop15 (Palmariggi-
Pop05 15 5.182 3.097 4.192 0.604 0.610 -0.036 1
Salento) was the smallest (FST = 0.0025; P ≤ 0.001); the dif-
Pop06 15 5.182 3.055 3.940 0.529 0.590 0.037 0
Pop07 16 5.636 3.213 4.156 0.557 0.544 -0.059 0 ferentiation coefficient between Pop02 (Deliceto-Daunian
Pop08 16 6.455 3.888 4.775 0.577 0.632 0.086* 0 sub-Apennine) and Pop22 (Vico del Gargano-East Gargano)
Pop09 16 7.182 4.166 4.903 0.605 0.709 0.155* 0 was the largest (FST = 0.1053; P < 0.01).
Pop10 16 6.455 3.979 4.601 0.587 0.656 0.093* 0 Nei’s genetic distance among the populations belonging
Pop11 16 6.000 3.369 4.313 0.665 0.614 -0.112 0 to the five units are graphically illustrated in Fig. 2, con-
Pop12 15 5.909 3.790 4.420 0.628 0.655 0.025 1 structed on the basis of principal coordinate analysis (PCoA
Pop13 12 5.545 3.883 4.760 0.595 0.705 0.092 2 with the physiographic units superimposed). The first two
Pop14 16 6.091 3.749 4.487 0.655 0.648 -0.018 1
axes of the PCoA explained 42.26% of the variation. The
Pop15 15 5.727 3.503 4.583 0.596 0.634 0.066 0
Pop16 11 5.273 3.400 4.606 0.627 0.647 0.006 1
scattergram did not show an evident correlation between
Pop17 13 6.091 3.700 4.560 0.580 0.639 0.062 0 the physiographic units and the genetic similarities of the
Pop18 14 5.909 3.688 4.651 0.605 0.642 0.005 1 populations. In fact, populations belonging to four differ-
Pop19 14 6.545 3.955 4.841 0.583 0.669 0.064 0 ent physiographic units are closely associated along the first
Pop20 14 7.091 4.363 5.171 0.617 0.692 0.120* 1 axis in the left part of the diagram. The Mantel test compar-
Pop21 16 6.000 3.753 4.510 0.557 0.684 0.116* 1 ing the matrices of geographic and genetic distances among
Pop22 11 4.727 2.559 3.941 0.527 0.540 -0.129 0 populations did not evidence any correlation between these
Mean 14.8 5.884 3.580 4.481 0.583 0.629 0.039 two variables.

Tab. 4. Analysis of molecular variance among populations, within populations, and within individuals of Quercus pubescens s.l. popula-
tions; Df – degree of freedom, SS – sum of squares, MS – mean squares, P – probability based on standard permutation across the full
data set, Nm –average estimate of gene flow among populations.
Source of variation Df SS MS Estimated variation Percentage variation % F-statistics P-value Gene flow
Among Populations 21 180.496 8.595 0.141 3.6 Fst = 0.036 Nm = 6.664
Among Individuals 303 1346.241 4.443 0.695 17.9 Fis = 0.185 0.001
Within Individuals 325 992.500 3.054 3.054 78.5 Fit = 0.215
Total 649 2519.237 3.889 100

ACTA BOT. CROAT. 79 (1), 2020 19


DI PIETRO R, DI MARZIO P, ANTONECCHIA G, CONTE AL, FORTINI P

netic clusters: a first cluster including 305 pure individu-


als (94.15%) and a second one composed of 19 off-type
individuals (5.85%) (Fig. 3a). The distribution of the 19
off-types involved 13 populations out of the 22 investigat-
ed (Fig. 3b, On-line Suppl. Tab. 4). Population 13, where
the sampled individuals were selected from a mixed wood
with Q. pubescens and Q. robur that developed in a humid
area, is the population showing the highest number (4) of
off-types. Populations 2, 9, and 10 count two off-types each
and relate to Q. pubescens s.l. woods with the occurrence
of Q. frainetto.
The Wilcoxon sign-rank test showed no significant re-
sults as regards bottleneck effect in all the populations an-
alyzed, using both TPM and SMM models. The observed
number of loci with heterozygosity excess (obs LHexc) was
found to be always lower than the expected number of loci
with heterozygosity excess (exp LHexc) except for popula-
Fig. 2. Principal Coordinate Analysis (PCoA) via Distance ma- tion 13 where “obs LHexc” is about 1.5 times greater than
trix (Nei Genetic Distance) with data standardization. Daunian “exp LHexc”. However, the Wilcoxon sign-rank test was not
sub-Apennine populations: Pop02, 06, 07, 08; W-Gargano popu- significant (Tab. 5).
lations: 04, 05; E-Gargano populations: 21, 22; Murgia plateau
populations: 01, 03, 09, 10, 12, 13, 14, 16, 17, 18, 19, 20; Salento
population: 15; Tavoliere population: 11. Discussion
The data-set exhibited the following levels of genetic di-
versity at the microsatellite loci examined: mean expected
When STRUCTURE was run using population IDs as heterozygosity (He) = 0.679, mean observed heterozygos-
sampling location information, a maximum value of the rate ity (Ho) = 0.583, average number of alleles per locus (K) =
of change in the log probability of data was revealed at K 11.0 (from 4 to 17 alleles). Four (PIE102, PIE215, PIE239,
= 2, using Evanno's method (Evanno et al. 2005) (On-line PIE267), out of the eleven loci investigated, showed a posi-
Suppl. Fig. 1). tive FIS (0.279-0417) which turned out to be statistically sig-
The individual membership proportion determined nificant (Tab. 2). It is known that the fixation of alleles could
by Bayesian clustering analysis suggested a subdivision of be favored by directional selection (Andolfatto 2001), or be
the whole data-set of oak individuals (325) into two ge- the results of genetic drift. Molecular studies regarding oth-

Fig. 3. Genetic assignment obtained with STRUCTURE clustering analysis. Oak individuals are distributed in progressive order along
the horizontal axis while the vertical axis expresses the group membership percentage degree per single individual per K=2. Clustering
of all the data-set oak individuals into two different genetic clusters with decreasing membership degrees for group 2 and increasing
degrees for group 1 (a). Clustering of all the data-set oak individuals ordered per populations; the latter separated by white lines (b).

20 ACTA BOT. CROAT. 79 (1), 2020


QUERCUS PUBESCENS COMPLEX IN SOUTHERN ITALY

Tab. 5. Results of tests for genetic bottlenecks using the two-phase the analysis of populations composed of a seemingly limited
mutation model (TPM) and a stepwise mutation model (SMM), number of samples, can still provide interesting information
and the Wilcoxon sign-rank tests for heterozygosity excess; that should not be completely ignored.
the number of polymorphic loci for all populations (Pop) is 11;
obs LHexc – observed number of loci with heterozygosity excess; The highest allelic variability was found in locus PIE152
exp LHexc – expected number of loci with heterozygosity excess; with 17 alleles and PIC = 0.879. A similar variability was
P – probability of no significant heterozygosity excess. found in other pubescent oak populations investigated in
TPM SMM southern Italy and central Italy (Antonecchia 2015) where
obs exp obs exp locus PIE152 exhibited the highest polymorphic degree and
Pop P P
LHexc LHexc LHexc LHexc in particular the highest values of allelic richness for Q. pu-
Pop01 4 6.46 0.9492 2 6.31 0.9954 bescens. The still insufficient knowledge about the Q. pube-
Pop02 6 6.44 0.7114 4 6.43 0.8398 scens genome do not allow it to be established if the high
Pop03 4 6.43 0.9385 4 6.43 0.9585
variability in locus PIE152 might be associated with adap-
Pop04 4 6.37 0.9126 4 6.37 0.9585
Pop05 3 6.29 0.8608 3 6.29 0.9263 tive traits. More in-depth investigations would be needed to
Pop06 5 6.39 0.7676 4 6.35 0.8799 clarify this point.
Pop07 4 6.36 0.9585 4 6.38 0.9731 The level of genetic variability within the pubescent oak
Pop08 5 6.49 0.9492 4 6.58 0.9663
Pop09 3 6.55 0.9492 2 6.50 0.9976
populations of the Apulia region was here found to be low-
Pop10 5 6.27 0.7676 3 6.36 0.8970 er than that reported for Q. pubescens in other countries.
Pop11 2 6.39 0.9895 2 6.49 0.9966 Curtu et al. (2007b) reported aHe = 0.891, Ar = 17.8 on a
Pop12 5 6.49 0.5845 5 6.57 0.7114 study based on 6 microsatellite loci for 73 Q. pubescens in-
Pop13 8 6.52 0.1602 8 6.61 0.2598 dividuals while Enescu et al. (2013) reported aHe=0.847, k
Pop14 4 6.59 0.8608 4 6.58 0.9585
Pop15 6 6.58 0.7676 5 6.51 0.8608 = 22 and Ar = 14.22 in a 7 microsatellite loci study in a geo-
Pop16 4 6.59 0.9263 4 6.55 0.9663 graphically related area. The heterozygosity values reported
Pop17 5 6.46 0.8799 2 6.51 0.9919 in these studies, however, would not be directly comparable
Pop18 3 6.47 0.9126 3 6.50 0.9126 with those obtained by us due to the use of different type of
Pop19 4 6.39 0.9126 4 6.50 0.9492 markers (genomic SSRs vs. EST-SSRs). Some authors (Du-
Pop20 4 6.48 0.8799 3 6.59 0.9919
Pop21 6 6.56 0.6812 4 6.57 0.8970 rand et al. 2010, Parthiban et al. 2018) showed that the use
Pop22 3 6.56 0.9663 2 6.53 0.9954 of EST-SSRs highlights the values of polymorphism slightly
lower than those obtained from the use of genomic SSRs.
However, there are some studies on the Quercus genus in
which both these types of markers were tested. Curtu et al.
er oak species stated that the genetic variation removed by
(2011) investigated the genetic variability of 65 individuals of
genetic drift would affect the genome rather uniformly (Al-
Q. pubescens using 7 microsatellite loci, of which 5 gSSR and
berto et al. 2010). This would not seem to be the case of the
2 EST-SSRs. The overall average values for the genetic pa-
Apulian pubescent oak populations, since the other EST-SSR
rameters in issue were the following: aHe = 0.859, k = 17, Ar
loci examined showed only slightly positive values of FIS, or
= 16.5. Calculating the values of these parameters solely for
even negative ones. Due to the relatively low number of sam-
the two EST-SSRs markers we obtained values (aHe = 0.819,
pled individuals per population, it cannot be completely ex-
k = 17, Ar = 16.8) that are lower than those obtained aver-
cluded that these results could be addressed, at least in part,
aging both kind of markers but still significantly higher than
to the sampling effect. However, there is no uniformity of
those we obtained from our Apulian pubescent oak data-set.
views at present on the minimum number of individuals per
population that should be sampled to detect “reliable” results The only available analyses of genetic differentiation
for genetic differentiation. Hale et al. (2012) stated that the within Italian white oak populations including Q. pubescens
number of individuals per population should be established s.l. carried out using EST-SSRs markers are those reported
using a “case by case” approach. On the other hand the ge- for the mixed Q. petraea, Q. frainetto, Q. pubescens woods of
netic literature on this issue shows a wide range of opinions the Mount Vairano range in southern Italy (Antonecchia et
and experiences from which it emerges that the number of al. 2015). Comparison between these analyses and our data
individuals sampled can be highly variable. In many cases revealed that the genetic variability of the Apulian popula-
it was considered as acceptable to take a minimum of 10 to tions is significantly lower than that recorded in Mount Vai-
30 individuals. For some Pinus sylvestris L. populations in rano (aHe = 0.72, k = 9.46 Ar = 9.2) as regards aHe and Ar,
the Baltic area, the number of 20-25 individuals was found and slightly higher as regards k.
to be large enough to detect all the alleles (Danusevičius et Also the level of genetic variability among the Q. pube-
al. 2016) whereas other studies have even considered sig- scens s.l. populations of the Apulia region (FST = 0.036) was
nificantly lower numbers (Kitamura et al. 2017, Rinaldi et found to be lower than the FST values found for other oak
al. 2019). In contrast, in Kalinowski (2005) it was shown species in other parts of the world, e.g. Q. rubra L. (FST =
particularly clearly that when FST is less than 0.01 it can be 0.080) (Sullivan et al. 2013), Quercus mongolica Fisch. ex
useful to sample up to 100 individuals. As far as we are con- Ladeb. (FST = 0.077) (Ueno and Tsumura 2008), Q. variabi-
cerned, we believe that these results, although deriving from lis Bl. (FST = 0.063) (Shi et al. 2017).
ACTA BOT. CROAT. 79 (1), 2020 21
DI PIETRO R, DI MARZIO P, ANTONECCHIA G, CONTE AL, FORTINI P

As regards gene flow, the present study showed that the As regards possible taxonomic implications arising from
Apulian Q. pubescens s.l. populations are characterized by a this study, the results of STRUCTURE Bayesian clustering
relatively high values (Nm = 6.664). Although Q. pubescens is showed that the populations of Apulian pubescent oaks
a typical wind-pollinated and outcrossing species such high could not be divided into groups. None of the populations
values are quite surprising considering the progressive con- investigated displayed significant differences in their genetic
traction suffered by the Apulian oak woods in the last mil- composition. However, the highest percentage of the possi-
lennium. The value of gene flow is determined by many fac- ble “putative hybrids” that we simply named “off-types”, was
tors, such as the biological means of spreading via pollen and found in Pop13, which also showed the highest percentage of
seeds, the occurrence of physical barriers among populations, private alleles. Pop13 is located within a flat sub-humid area
population dimension and so forth; in addition, it can be fa- in the western side of the Murgian plateau in the municipal-
cilitated by physical proximity of the populations. It is prob- ity of Laterza, and lies in spatial contact with a small stand
able that the short distance separating some of the Apulian of Q. robur wood (the only known site of Q. robur currently
pubescent oak populations and the lack of significant moun- recorded for the Apulia region). It is conceivable that in this
tain barriers have contributed to keep the average gene flow area Q. robur was much more abundant in the past and that
value high. It is not possible, at present, to establish whether both climate change and the millennial work of deforesta-
the above gene-flow values are stable over time, or are the tion carried out by the local populations, has led to its cur-
result of a contraction due to over-exploitation of forests. In rent extreme impoverishment. It is very probable that during
studies on other oak species (Shi et al. 2017) it was established the climatic oscillations that took place during the Quater-
that gene-flow values can decrease significantly in a relatively nary, environmental situations favorable to the development
short time (ten years) following the transition from primary of mixed forests of Q. robur and Q. pubescens s.l. were cre-
forest to natural secondary forest. In the Apulia region the ated several times with consequent hybridization or intro-
millennial exploitation of the woodlands led to the disappear- gression between these two species. In fact, a possible key of
ance of primeval forests as early as Roman times. However, interpretation of genetic structure of the Q. pubescens Apu-
the secondary forest communities that have since replaced lian populations cannot fail to consider some physiographic
the primary ones show a high degree of naturalness and a and geographical features of this region such as the lack of
rather stable floristic composition. It is conceivable therefore sizeable mountain systems and its deep wedges jutting into
that the gene flow rate will remain stable over time and that the Adriatic Sea. These features led to the Apulian Peninsula
the only parameter capable of negatively acting on the gene- experiencing the effects of quaternary glaciations only mild-
flow rate is the spatial reduction of the population size. ly and allowed some parts of it to work as glacial refuges for
The apparent relatively low level of intraspecific diversi- the thermophilous oak forests during the Quaternary cold
ty observed for the Apulian Q. pubescens populations could periods, when most of the Italian Peninsula was covered by
be addressed to their marginal geographical position not steppic grasslands (Follieri et al. 1988). The occurrence of
only in the Italian Peninsula but also in the central Med- restricted areas where the climate forced the various species
iterranean context. In terms of biological conservation of of white oaks to coexist for long (cold) periods may have fa-
deciduous oak forests the Apulia region is particularly ex- vored the generation of hybrids. The innumerable past hy-
posed to environmental risks due to climate change drying bridization or introgression events, which led to the current
effects which overlap a bioclimatic pattern dominated by the high morphological variability of Q. pubescens s.l., may also
Thermo-Mediterranean thermotype the latter having in the have played a role in enlarging the gene flow of the Apulian
Mediterranean maquis the main type of potential vegeta- populations. White oaks commonly live in sympatry, mak-
tion (Rivas-Martínez et al. 2004, Ladisa et al. 2012). Threats ing them particularly suitable for gene exchange and pro-
to pubescent oak woods are further increasing in many lo- duction of hybrid individuals (Rushton 1993, Williams et al.
calities that are experiencing greater aridity and water defi- 2001, Lepais and Gerber 2011). According to Burger (1975),
ciency due to the overexploitation of water reserves result- applying the biological concept of species to the genus Quer-
ing from the ever increasing consumption for agricultural cus would cause an upheaval of the nomenclature, since the
purposes and by tourist settlements. In the long term, the binomial species would then not correspond to the biologi-
combined effect of these two factors could lead to a change cal one. This makes the systematics of this genus extremely
in the floristic composition of the upper structural layers of susceptible to disagreement among botanists and the hope
the thermophilous deciduous oak forests with a progressive of arriving at the proposal of a largely shared and possibly
replacement of the pubescent oak individuals by evergreen unambiguous taxonomic framework (in particular for the
woody species, such as Quercus ilex L. Phillyrea latifolia L., south-European white oaks) is something not soon to be re-
Rhamnus alaternus L., Arbutus unedo L. and Pistacia lentis- alized. The problematic classification of pubescent oaks in
cus L. Conversely, the high gene flow values observed in the the Apulian Peninsula and in the entire Italian Peninsula is
study area seem to move in the opposite direction and might perfectly evidenced by the discrepancy that emerges when
positively contribute to preventing possible activations of ge- comparing the taxonomy of the genus Quercus as reported in
netic drift. Yet the lack of bottleneck effects in all investigated latest versions of Flora d’Italia (Pignatti et al. 2017) and in the
populations implies a certain degree of resilience despite the Checklist of the Italian vascular flora (Bartolucci et al. 2018).
relatively small areas covered by the single oak populations. Although both works are worthy of the utmost respect and
22 ACTA BOT. CROAT. 79 (1), 2020
QUERCUS PUBESCENS COMPLEX IN SOUTHERN ITALY

consideration, they show enormous differences in both the in south-eastern Europe, such as the eternal debate on the
numbers and the names of the taxa considered good species dualism between Q. pubescens and Q. virgiliana (cf. Borazan
for the same taxonomic group (white oaks). Actually, based and Babaç 2003, Škvorc et al. 2005, Franjić et al. 2006, Jerše
on what emerges from our work, the high taxonomic dif- and Batič 2007, Trinajstić 2007, Enescu et al. 2013). We sim-
ferentiation among the Italian pubescent oaks as reported ply believe that this work represents a small step forward, a
in most of the Italian floras and checklists does not seem to substantial contribution, albeit on a regional scale, to being
have much foundation (see also Di Pietro et al. 2012) or at able to arrive in the near future at a possible solution of the
least does not seem to have it with regard to the pubescent problem on a global scale. Finally, it is worth emphasizing,
oaks of Apulian Peninsula. In fact, our results established that all the taxonomical and phytosociological considerations
that it was not possible to identify genetic clusters among mentioned in this paragraph should not be viewed as useful
the pubescent oak populations of the Apulian region basing only for updating floras, checklists or vegetation maps. In
on 11 highly polymorphic markers as it was not possible to fact, they have direct implications for the application of Eu-
identify morphological clusters basing in a previous biomet- ropean Policies for Nature conservation. It is only necessary
ric study on the morphological traits of leaves and fruits on to recall the 92/43/EEC European directive where the identi-
the same set of specimens (Di Pietro et al. 2016). fication of some of the forest habitats listed in the Manual of
Summarizing, the presence of more than one taxon at Interpretation (e.g., 91H0* “Pannonian woods with Quercus
species rank belonging to Q. pubescens s.l. (e.g., Q. virgiliana, pubescens”, 91AA* “Eastern white oak woods”, 91M0 “Panno-
Q. dalechampii, Q. amplifolia, etc.) is not confirmed (at least at nian-Balkanic turkey oak-sessile oak forests”, 91Y0 “Dacian
present) for the Apulia region. At the same time, the phytoso- oak and hornbeam forests”), is also based on the occurrence
ciological frameworks that are currently supposed to classify of species such as Q. virgiliana and Q. dalechampii.
different types of pubescent oak forests (phytosociological
associations), dominated by different pubescent oak species Acknowledgments
and including up to five different species of pubescent oaks
in the dominant tree layer, need at the very least to be recon- The authors wish to acknowledge G. Vendramin for his
sidered. Obviously we do not consider this work to be a so- critical discussion of the preliminary results of the paper and
lution to all the problems concerning the taxonomy of white P. Medagli, G. Misano, G. Silletti, V. Viscosi, R.P. Wagensom-
pubescent oaks, not even to some of them of topical relevance mer for their help in collecting oak populations.

References
Alberto, F., Niort, J., Derory, J., Lepais, O., Vitalis, R., Galop, D., tangelo, A., Scoppola, A., Scortegagna, S., Selvaggi, A., Selvi,
Kremer, A., 2010: Population differentiation of sessile oak at F., Soldano, A., Stinca, A., Wagensommer, R.P., Wilhalm, T.,
the altitudinal front of migration in the French Pyrenees. Mo- Conti, F., 2018: An updated checklist of the vascular flora na-
lecular Ecology 19, 2626–2639. tive to Italy. Plant Biosystems 152, 179–303.
Andolfatto, P., 2001: Adaptive hitchhiking effects on genome vari- Biondi, E., Casavecchia, S., Beccarisi, L., Marchiori, S., Medagli,
ability. Current Opinion in Genetics and Development 11, P., Zuccarello, V., 2010: Le serie di vegetazione della Regione
635–641. Puglia. In: Blasi, C. (ed.), La vegetazione d’Italia, 391–409. Pa-
Antonecchia, G., 2015: Analisi della variabilità genetica del sot- lombi Editori, Roma.
togenere Quercus Oerst. in Italia centro-meridionale. PhD Biondi, E., Casavecchia, S., Guerra, V., Medagli, P., Beccarisi, L.,
thesis, Department of Bioscience and Territory, University of Zuccarello, V., 2004: A Contribution towards the knowledge
Molise, Pesche (IS). Retrieved April 10, 2016 from http://hdl. of semideciduous and evergreen woods of Apulia (southeast-
handle.net/2192/296. ern Italy). Fitosociologia 41, 3–28.
Antonecchia, G., Fortini, P., Lepais, O., Gerber, S., Léger, P., Scip- Biondi, E., Guerra, V., 2008: Vegetazione e paesaggio vegetale del-
pa, G. S., Viscosi, V., 2015: Genetic structure of a natural oak le gravine dell'arco jonico. Fitosociologia 45, 57–125.
community in central Italy, Evidence of gene flow between Blasi, C., Michetti, L., 2007: Biodiversity and climate. In: Blasi, C.,
three sympatric white oak species (Quercus, Fagaceae). An- Boitani, L., La Posta, S., Manes, F., Marchetti, M. (eds.), Bio-
nals of Forest Research 58, 205–216. diversity in Italy. Contribution to the National Biodiversity
Ballian, D., Ivanković, M., Gračan, J., Perić, S., Marjanović, H., Bo- Strategy, 57–66. Palombi Editori, Roma.
binac, M., Slade, D., 2010: Analysis of pubescent oak (Quercus Borazan, A., Babaç, M.T., 2003: Morphometric leaf variation in
pubescens Willd.) in the western part of the Balkan peninsula. oaks (Quercus) of Bulu, Turkey. Annales Botanici Fennici 40,
Acta Societatis Botanicorum Poloniae 79, 189–195. 233–242.
Bartolucci, F., Peruzzi, L., Galasso, G., Albano, A., Alessandrini, Burger, W.C., 1975: The Species Concept in Quercus. Taxon 24,
A., Ardenghi, N.M.G., Astuti, G., Bacchetta, G., Ballelli, S., 45–50.
Banfi, E., Barberis, G., Bernardo, L., Bouvet, D., Bovio, M., Cornuet, J.M., Luikart, G., 1996: Description and power analysis
Cecchi, L., Di Pietro, R., Domina, G., Fascetti, S., Fenu, G., of two tests for detecting recent population bottlenecks from
Festi, F., Foggi, B., Gallo, L., Gottschlich, G., Gubellini, L., Ia- allele frequency data. Genetics 144, 2001–2014.
monico, D., Iberite, M., Jiménez‐Mejías, P., Lattanzi, E., Mar- Cotecchia, V., Simeone, V., Gabriele S., 2014: Caratteri climati-
chetti, D., Martinetto, E., Masin, R.R., Medagli, P., Passalac- ci della Regione Puglia. In: Ispra Ambiente (ed.), Memorie
qua, N.G., Peccenini, S., Pennesi, R., Pierini, B., Poldini, L., descrittive della Carta geologica d’Italia 92, 338–369. ISPRA
Prosser, F., Raimondo, F.M., Roma‐Marzio, F., Rosati, L., San- serv. Geologico d’Italia, Roma.

ACTA BOT. CROAT. 79 (1), 2020 23


DI PIETRO R, DI MARZIO P, ANTONECCHIA G, CONTE AL, FORTINI P

Curtu, A.L., Gailing, O., Finkeldey, R., 2007a: Evidence for hybrid- scens (Fagaceae), insights from macro-morphological leaf
ization and introgression within a species-rich oak (Quercus traits and molecular data. Plant Systematics and Evolution
spp.) community. BMC Evolutionary Biology 7, 218. 301, 375–385.
Curtu, A.L., Gailing, O., Finkeldey, R., 2009: Patterns of contem- Fortini, P., Viscosi, V., Maiuro, L., Fineschi, S., Vendramin, G.G.,
porary hybridization inferred from paternity analysis in a 2009: Comparative leaf surface morphology and molecular
four-species forest. BMC Evolutionary Biology 9: 284. data of five oaks of the subgenus Quercus Oerst (Fagacee).
Curtu, A.L., Gailing, O., Leinemann, L., Finkeldey, R., 2007b: Ge- Plant Biosystems 143, 543–554.
netic variation and differentiation within a natural communi- Franjić, J., Liber, B., Škvorc, Z., Idžojtić, M., Šostarić, R., Stančić,
ty of five oak species (Quercus spp.). Plant Biology 9, 116–126. Z., 2006: Morphological and molecular differentiation of the
Curtu, A.L., Moldovan, I.C., Enescu, C.M., Craciunesc, J., So- croatian populations of Quercus pubescens Willd. (Fagaceae).
fletea, N., 2011: Genetic Differentiation between Quercus Acta Societatis Botanicorum Poloniae 75, 123–130.
frainetto Ten. and Q. pubescens Willd. in Romania. Notulae Gailing, O., Bodénès, C., Finkeldey, R., Kremer, A., Plomion, C.,
Botanicae Horti Agrobotanici Cluj-Napoca 39, 275–282. 2013: Genetic mapping of EST-derived Simple Sequence Re-
Danusevičius, D., Kavaliauskas, D., Fussi, B., 2016: Optimum peats (EST-SSRs) to identify QTL for leaf morphological char-
acters in a Quercus robur full-sib family. Tree Genetics and
sample size for SSR-based estimation of representative allele
Genomes 9, 1361–1367.
frequencies and genetic diversity in Scots pine populations.
Baltic Forestry 22, 194–202. Goudet, J., 2003: FSTAT (version 2.9.4), a program (for Windows
95 and above) to estimate and test population genetics param-
Di Pietro, R., Di Marzio, P., Medagli, P., Misano, G., Silletti, G.N.,
eters. Retrieved November 26, 2015 from http://www2.unil.
Wagensommer, R.P., Fortini, P., 2016: Evidence from multi- ch/izea/softwares/fstat.html
variate morphometric study of the Quercus pubescens com-
Govaerts, R., Frodin, D.G., 1998: World checklist and bibliogra-
plex in southeast Italy. Botanica Serbica 40, 83–100.
phy of Fagales (Betulaceae, Corylaceae, Fagaceae and Tico-
Di Pietro, R., Misano, G., 2009: Analisi fitosociologica e conside- dendraceae). Royal Botanic Gardens: Kew, Richmond.
razioni sintassonomiche sulla vegetazione forestale delle Gra-
Guichoux, E., Lagache, L., Wagner, S., Chaumeil, P., Léger, P., Le-
vine occidentali dell’Arco Ionico (Murge pugliesi e lucane, Ita-
pais, O., Lepoittevein, C., Malausa, T., Revardel, E., Salin, F.,
lia meridionale). Informatore Botanico Italiano 41, 215–246. Petit, R.J., 2011: Current trends in microsatellite genotyping.
Di Pietro, R., Viscosi, V., Peruzzi, L., Fortini, P., 2012: A review Molecular Ecology Resources 11, 591–611.
of the application of the name Quercus dalechampii. Taxon Hale, M.L., Burg, T.M., Steeves, T.E., 2012: Sampling for micro-
61, 1311–1316. satellite-based population genetic studies: 25 to 30 individu-
Durand, J., Bodénès, C., Chancerel, E., Frigerio, Vendramin, als per population is enough to accurately estimate allele fre-
G., Sebastiani, F., Buonamici, A., Gailing, O., Koelewijn, quencies. PLOS ONE 7, e 45170.
H.P., Villani, F., Mattioni, C., Cherubini, M., Goicoechea, Hildebrand, C.E., Torney, D.C., Wagner, R.P., 1992: Informative-
P.B., Herrán, A., Ikaran, Z., Cabané, C., Ueno, S., Alberto, F., ness of polymorphic DNA markers. Los Alamos Science 20,
Dumoulin, P.Y., Guichoux, E., de Daruvar, A., Kremer, A., 100–102.
Plomion, C., 2010: A fast and cost-effective approach to de- Jerše, M., Batič, F., 2007: Morphological analysis of pubescent oak
velop and map ESTSSR markers, oak as a case study. BMC (Quercus pubescens Willd.) in Slovenia. Zbornik gozdarstva in
Genomics 11, 570. lesarstva 83, 35–45.
Earl, D.A., von Holdt, B.M., 2012: STRUCTURE HARVESTER, Kalinowski, S.T., 2005: Do polymorphic loci require large sample
a website and program for visualizing STRUCTURE output sizes to estimate genetic distances? Heredity 94, 33–36.
and implementing the Evanno method. Conservation Genet-
Kalinowski, S.T., Taper, M.L., Marshall, T.C., 2007: Revising how
ics Resources 4, 359–361.
the computer program CERVUS accommodates genotyping
Enescu, C.M., Curtu, A. L., Sofletea, N., 2013: Is Quercus virgili- error increases success in paternity assignment. Molecular
ana a distinct morphological and genetic entity among Eu- Ecology 16, 1099–1106.
ropean white oaks? Turkish Journal of Agriculture and For-
estry 37, 632–641. Kitamura, K., Namikawa, K., Kawahara, T., Matsumoto, A., San
Jose-Maldia, L., 2017: Genetic structure of remnant Quercus
Evanno, G., Regnaut, S., Goudet, J., 2005: Detecting the number
serrata populations at the northernmost limit of their distri-
of clusters of individuals using the software STRUCTURE, a
bution in Japan. Acta Phytotaxonomica et Geobotanica 68,
simulation study. Molecular Ecology 14, 2611–2620.
1–15.
Excoffier, L., Lischer, H.E.L., 2010: Arlequin suite ver 3.5: A new
series of programs to perform population genetics analyses Ladisa, G., Todorovic, M., Trisorio, L.G., 2012: A GIS-based ap-
under Linux and Windows. Molecular Ecology Resources 10, proach for desertification risk assessment in Apulia region, SE
564–567. Italy. Physics and Chemistry of the Earth 49, 103–113.
Fineschi, S., Taurchini, D., Grossoni, P., Petit, R.J., Vendramin, Lepais, O., Gerber, S., 2011: Reproductive patterns shape intro-
G.G., 2002: Chloroplast DNA variation of white oaks in Italy. gression dynamics and species succession within the Europe-
Forest Ecology and Management 156, 103–114. an white oak species complex. Evolution 65, 156–170.
Fineschi, S., Vendramin, G.G., 2004: La diversità cloroplastica del- Luikart, G., Cornuet, J.M., 1998: Empirical evaluation of a test for
le querce italiane: evidenze di una maggiore ricchezza geneti- identifying recently bottlenecked populations from allele fre-
ca nelle popolazioni meridionali e insulari. Forest@ 1, 82–87. quency data. Conservation Biology 12, 228–237.
Follieri, M., Magri, D., Sadori, L., 1988: 250,000-year pollen re- Nixon, K.C., 1993: Infrageneric classification of Quercus (Faga-
cord from Valle di Castiglione (Roma). Pollen et Spores 30, ceae) and typification of sectional names. Annals of Forest
329–256. Science 50 (Suppl. 1993), 25s–34s.
Fortini, P., Antonecchia, G., Di Marzio, P., Maiuro, L., Viscosi, V., Parthiban, S., Govindaraj, P., Senthilkumar, S., 2018: Comparison
2013: Role of micromorphological leaf traits and molecular of relative efficiency of genomic SSR and EST-SSR markers
data in taxonomy of three sympatric white oak species and in estimating genetic diversity in sugarcane. 3 Biotech 8, 144.
their hybrids (Quercus L.). Plant Biosystems 149, 546–558. Peakall, R., Smouse, P.E., 2012: GenAlEx 6.5, genetic analysis in
Fortini, P., Di Marzio, P., Di Pietro, R., 2015: Differentiation and Excel. Population genetic software for teaching and research-
hybridization of Quercus frainetto, Q. petraea, and Q. pube- an update. Bioinformatics 28, 2537–2539.

24 ACTA BOT. CROAT. 79 (1), 2020


QUERCUS PUBESCENS COMPLEX IN SOUTHERN ITALY

Peñuelas, J., Sardans, J., Filella, I., Estiarte, M., Llusià, J., Ogaya, R., China as revealed by microsatellites markers. Forests 8, 495.
Carnicer, J., Bartrons, M., Rivas-Ubach, A., Grau, O., Pegue- Škvorc, Z., Franjić, J., Idžojtić, M., 2005: Population structure
ro, G., Margalef, O., Pla-Rabés, S., Stefanescu, C., Asensio, D., of Quercus pubescens Willd. (Fagaceae) in Croatia accord-
Preece, C., Liu, L., Verger, A., Barbeta, A., Achotegui-Castells, ing to morphology of leaves. Acta Botanica Hungarica 47,
A., Gargallo-Garriga, A., Sperlich, D., Farré-Armengol, G., 193–206.
Fernández-Martínez, M., Liu, D., Zhang, C., Urbina, I., Cami- Slade, D., Škvorc, Ž., Ballian, D., Gračan, J., Papeš, D., 2008: The
no-Serrano, M., Vives-Ingla, M., Stocker, B.D., Balzarolo, M., chloroplast DNA polymorphisms of White Oaks of section
Guerrieri, R., Peaucelle, M., Marañón-Jiménez, S., Bórnez- Quercus in the Central Balkans. Silvae Genetica 57, 227–234.
Mejías, K., Mu, Z., Descals, A., Castellanos, A., Terradas, J.,
2017: Impacts of global change on Mediterranean forests and Slatkin, M., Barton, N.H., 1989: A comparison of three indirect
their services. Forests 8, 463. methods for estimating average levels of gene flow. Evolution
43, 1349–1368.
Pignatti, S., Guarino, R., La Rosa, M., 2017: Flora d’Italia. New
Business Media, Bologna. Sullivan, A.R., Lind, J.F., McCleary, T.S., Romero-Severson, J., Gail-
ing, O., 2013: Development and characterization of genomic
Piry, S., Luikart, G., Cornuet, J.-M., 1999: Computer note. BOT-
and gene-based microsatellite markers in north American red
TLENECK, a computer program for detecting recent reduc-
tions in the effective size using allele frequency data. The Jour- oak species. Plant Molecular Biology Reporter 31, 231–239.
nal of Heredity 90, 502–503. Toonen, R.J., Hughes, S., 2001: Increased throughput for fragment
Pritchard, J.K., Stephens, M., Donnelly, P., 2000: Inference of pop- analysis on an ABI PRISM 377 automated sequencer using a
ulation structure using multilocus genotype data. Genetics membrane comb and STRand software. Biotechniques 31,
155, 945–959. 1320–1324.
Pritchard, J.K., Wen, X., Falush, D., 2010: Documentation for Trinajstić, I., 2007: About the problem of differentiation between
STRUCTURE software. Retrieved March 1, 2019 from http:// the oaks Quercus pubescens Willd and Quercus virgiliana
pritch.bsd.uchicago.edu/structure.html (Ten.). Šumarski list 131, 57–60.
Rinaldi, R., Cafasso, D., Strumia, S., Cristaudo, A., Sebastiani, F., Ueno, S., Tsumura, Y., 2008: Development of ten microsatellite
Fineschi, S., 2019: The influence of a relict distribution on ge- markers for Quercus mongolica var. crispula by database min-
netic structure and variation in the Mediterranean tree, Plata- ing. Conservation Genetics 9, 1083–1085.
nus orientalis. AoB PLANTS 11: plz002. Van Oosterhout, C., Hutchinson, W.F., Wills, D.P.M., Shipley, P.,
Rivas-Martínez, S., Penas, A., Díaz, T.E., 2004: Bioclimatic and 2004: MICRO‐CHECKER, software for identifying and cor-
biogeographic maps of Europe. University of León. Retrieved recting genotyping errors in microsatellite data. Molecular
March 15, 2018 from www.globalbioclimatics.org/form/ Ecology Notes 4, 535–538.
maps.htm Weir, B.S., Cockerham, C.C., 1984: Estimating F‐statistics for the
Rushton, B.S., 1993: Natural hybridization within the genus Quer- analysis of population structure. Evolution 38, 1358–1370.
cus L. Annals of Forest Science 50 (Suppl. 1993), 73s–90s. Williams, J.H., Boecklen, W.J., Howard, D.J., 2001: Reproductive
Shi, X., Wen, Q., Cao, M., Guo, X., Xu, L., 2017: Genetic diver- processes in two oak (Quercus) contact zones with different
sity and structure of natural Quercus variabilis population in levels of hybridization. Heredity 87, 680–690.

ACTA BOT. CROAT. 79 (1), 2020 25

You might also like